首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 52 毫秒
1.
The evolution of envelope mutations by replicating primate immunodeficiency viruses allows these viruses to escape from the immune pressure mediated by neutralizing antibodies. Vaccine-induced anti-envelope antibody responses may accelerate and/or alter the specificity of the antibodies, thus shaping the evolution of envelope mutations in the replicating virus. To explore this possibility, we studied the neutralizing antibody response and the envelope sequences in rhesus monkeys vaccinated with either gag-pol-nef immunogens or gag-pol-nef immunogens in combination with env and then infected with simian immunodeficiency virus (SIV). Using a pseudovirion neutralization assay, we demonstrate that envelope vaccination primed for an accelerated neutralizing antibody response following virus challenge. To monitor viral envelope evolution in these two cohorts of monkeys, full-length envelopes from plasma virus isolated at weeks 37 and 62 postchallenge were sequenced by single genome amplification to identify sites of envelope mutations. We show that env vaccination was associated with a change in the pattern of envelope mutations. Prevalent mutations in sequences from gag-pol-nef vaccinees included deletions in both variable regions 1 and 4 (V1 and V4), whereas deletions in the env vaccinees occurred only in V1. These data show that env vaccination altered the focus of the antibody-mediated selection pressure on the evolution of envelope following SIV challenge.Immune containment of human immunodeficiency virus (HIV-1) is complicated by the continuous genetic evolution of the virus. The evolution of the HIV-1 envelope is shaped, in part, by selective pressure of neutralizing antibodies (6, 12, 27, 34-36, 40). Changes in envelope sequence and glycosylation patterns following infection can allow the virus to escape neutralization. If the rate and extent of envelope sequence evolution following infection can be decreased, immune containment of HIV-1 may be improved.One possible strategy for modifying envelope evolution is vaccination prior to infection. A vaccine-elicited memory immune response could focus and potentiate the humoral immune response that develops following infection. The possible consequence of vaccination has not been assessed, however, because of the limited number of human volunteers who have received highly immunogenic envelope immunogens and subsequently became infected with HIV-1.Simian immunodeficiency virus (SIV) infection of rhesus monkeys provides a powerful model to study the effect of vaccination on envelope evolution. Like HIV-1, SIV employs both the CD4 molecule and the chemokine receptor CCR5 to enter a target cell and cause an AIDS-like disease in macaques (16, 22). Both SIV and HIV-1 envelopes are heavily glycosylated, with approximately 50% of their mass derived from carbohydrates (14, 21). SIV and HIV-1 envelopes share approximately 40% amino acid homology (10, 11) and have overlapping variable and constant regions, although the variable region 3 (V3) of HIV-1 envelope does not align with the homologous region of SIV envelope (7). Following SIV infection in rhesus monkeys, SIV envelope evolves most rapidly in variable regions 1 and 4 (V1 and V4, respectively), leading to nucleotide additions, deletions, and/or mutations that can potentially translate to changes in glycosylation (7, 9, 13, 15, 19, 29, 30).Studies done to characterize SIV neutralization suggest that it occurs through mechanisms similar to those seen in HIV-1 neutralization. Amino acid mutations in the envelope of both viruses contribute to the evasion of antibody binding directly by changing recognition sequences and/or envelope conformation. In addition, the glycosylation of envelope serves as a further obstacle to antibody recognition (20, 33, 40). Considerable effort has been devoted to defining neutralizing epitopes of the HIV and SIV envelopes. The known neutralizing human monoclonal antibodies elicited during natural infection are directed against HIV-1 envelope target sites on both gp120 and gp41, including the V3 region, the CD4 binding site, oligomannose residues of gp120, and gp41 (17, 31). The neutralizing epitope profile of SIV envelope includes the CD4 binding site and gp41 but not the V3 region. There is conflicting evidence as to whether V1, V2, and/or V4 of SIV are targets for antibody neutralization (15, 18, 19). The present study addresses whether vaccine-induced immune responses accelerate the generation of autologous neutralizing antibodies following SIV challenge in rhesus monkeys and how this humoral immune response can potentially shape viral sequence evolution.  相似文献   

2.
Hepatitis C virus (HCV) envelope glycoproteins are highly glycosylated, with generally 4 and 11 N-linked glycans on E1 and E2, respectively. Studies using mutated recombinant HCV envelope glycoproteins incorporated into retroviral pseudoparticles (HCVpp) suggest that some glycans play a role in protein folding, virus entry, and protection against neutralization. The development of a cell culture system producing infectious particles (HCVcc) in hepatoma cells provides an opportunity to characterize the role of these glycans in the context of authentic infectious virions. Here, we used HCVcc in which point mutations were engineered at N-linked glycosylation sites to determine the role of these glycans in the functions of HCV envelope proteins. The mutants were characterized for their effects on virus replication and envelope protein expression as well as on viral particle secretion, infectivity, and sensitivity to neutralizing antibodies. Our results indicate that several glycans play an important role in HCVcc assembly and/or infectivity. Furthermore, our data demonstrate that at least five glycans on E2 (denoted E2N1, E2N2, E2N4, E2N6, and E2N11) strongly reduce the sensitivity of HCVcc to antibody neutralization, with four of them surrounding the CD81 binding site. Altogether, these data indicate that the glycans associated with HCV envelope glycoproteins play roles at different steps of the viral life cycle. They also highlight differences in the effects of glycosylation mutations between the HCVpp and HCVcc systems. Furthermore, these carbohydrates form a “glycan shield” at the surface of the virion, which contributes to the evasion of HCV from the humoral immune response.Hepatitis C virus (HCV) is a single-stranded positive-sense RNA virus that causes serious liver diseases in humans (31). More than 170 million people worldwide are seropositive for HCV and at risk for developing cirrhosis and hepatocellular carcinoma (50). HCV is a small, enveloped virus that belongs to the Hepacivirus genus in the Flaviviridae family (31). Its genome encodes a single polyprotein precursor of about 3,000-amino-acid residues that is cleaved co- and posttranslationally by cellular and viral proteases to yield at least 10 mature products (31). The two envelope glycoproteins, E1 and E2, are released from the polyprotein by signal peptidase cleavages. These two proteins assemble as noncovalent heterodimers, which are retained mainly in the endoplasmic reticulum (ER) (36), and they are found as large disulfide-linked oligomers on the surfaces of HCV particles (46). HCV glycoproteins are involved in the entry process, and since they are present on the surfaces of viral particles, these proteins are the targets of neutralizing antibodies (4, 21).E1 and E2 generally contain 4 and 11 N-glycosylation sites, respectively, all of which have been shown to be modified by glycans (19). Despite variability in HCV envelope glycoprotein sequences, the four glycosylation sites of E1 and nine of E2 are highly conserved, suggesting that the glycans associated with these proteins play an essential role in the HCV life cycle (22). Using retroviral particles pseudotyped with genotype 1a (H strain) HCV envelope glycoproteins (HCVpp), recent studies have determined the potential roles played by these glycans in protein folding, HCV entry, and protection against neutralization (14, 19, 22). Indeed, the lack of glycan E1N1, E1N4, E2N8, or E2N10 strongly affects the incorporation of HCV glycoproteins into HCVpp, suggesting that these glycans are important for correct protein folding (19). Furthermore, mutation of glycosylation sites E2N2 or E2N4 alters HCVpp infectivity despite normal incorporation into pseudotyped particles, suggesting a role for the corresponding glycans in viral entry, at least in this model system (19). Finally, glycans at positions E2N1, E2N6, and E2N11 were shown to reduce the sensitivity of HCVpp to antibody neutralization as well as access of the CD81 coreceptor to its binding site on E2, suggesting that glycans also contribute to HCV evasion of the humoral immune response (14, 22).It has recently been proposed that targeting glycans could be a promising approach to inhibiting viral infection (1). Indeed, HCV, as well as several other viruses with highly glycosylated envelope proteins, can be inhibited by carbohydrate binding agents such as cyanovirin-N and pradimicin A (1, 7, 23). Furthermore, resistance against drugs that target glycans is likely to develop and will probably result in mutations at some glycosylation sites (3, 52). However, since glycans associated with viral envelope proteins play an important role in the viral life cycle, adaptation of viruses to the selective pressure of carbohydrate-binding agents will most likely come at a replicative cost to the virus (2).Although the role of HCV glycans has been studied using mutant recombinant HCV envelope glycoproteins incorporated into HCVpp, these particles do not recapitulate all the functions of HCV envelope proteins. Cell culture-derived virus (HCVcc) (32, 49, 55) assembles in an ER-derived compartment in association with very low density lipoproteins (17, 26), whereas HCVpp are assembled in a post-Golgi compartment and are not associated with lipoproteins (44). Importantly, this leads to differences between HCVpp and HCVcc in the oligomerization of the envelope glycoproteins (46). It is also important to note that the carbohydrate composition of viral glycoproteins can differ when the same virus is grown in different cell lines (13). Thus, HCVpp that are produced in 293T cells are not the most appropriate model for glycosylation studies, since HCV tropism is restricted to the liver. Furthermore, differences in envelope protein glycosylation have been observed between HCVpp and HCVcc particles (46). Differences in some HCV envelope protein functions were also observed when the HCVpp and HCVcc systems were compared (28, 29, 42, 43). The development of the HCVcc system provides, therefore, the opportunity to characterize the role of E1/E2-associated glycans in the context of authentic infectious virions. Here, we analyzed the role of E1/E2 glycans by introducing point mutations at N-linked glycosylation sites in the context of the HCVcc system. The effects of these mutations on virus replication, particle secretion, infectivity, and sensitivity to neutralizing antibodies were investigated. Our results demonstrate that several glycans play an important role in HCVcc assembly and/or infectivity and reduce access of neutralizing antibodies to their epitopes.  相似文献   

3.
4.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  相似文献   

5.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

6.
A major obstacle to gene transduction by viral vectors is inactivation by human complement in vivo. One way to overcome this is to incorporate complement regulatory proteins, such as CD55/decay accelerating factor (DAF), into viral particles. Lentivirus vectors pseudotyped with the baculovirus envelope protein GP64 have been shown to acquire more potent resistance to serum inactivation and longer transgene expression than those pseudotyped with the vesicular stomatitis virus (VSV) envelope protein G. However, the molecular mechanisms underlying resistance to serum inactivation in pseudotype particles bearing the GP64 have not been precisely elucidated. In this study, we generated pseudotype and recombinant VSVs bearing the GP64. Recombinant VSVs generated in human cell lines exhibited the incorporation of human DAF in viral particles and were resistant to serum inactivation, whereas those generated in insect cells exhibited no incorporation of human DAF and were sensitive to complement inactivation. The GP64 and human DAF were detected on the detergent-resistant membrane and were coprecipitated by immunoprecipitation analysis. A pseudotype VSV bearing GP64 produced in human DAF knockdown cells reduced resistance to serum inactivation. In contrast, recombinant baculoviruses generated in insect cells expressing human DAF or carrying the human DAF gene exhibited resistance to complement inactivation. These results suggest that the incorporation of human DAF into viral particles by interacting with baculovirus GP64 is involved in the acquisition of resistance to serum inactivation.Gene therapy is a potential treatment option for genetic diseases, malignant diseases, and other acquired diseases. To this end, safe and efficient gene transfer into specific target cells is a central requirement, and a variety of nonviral and viral vector systems have been developed (6, 44). Recombinant viruses can be used for efficient gene transfer. Retroviruses, adeno-associated viruses, and lentiviruses are able to integrate foreign genes into host genomes and are suitable for gene therapeutics by virtue of their permanent expression of the therapeutic genes, whereas adenoviruses, herpesviruses, and baculoviruses can transiently express foreign genes (6, 12, 44). Pseudotype particles bearing other viral envelope proteins have been developed to improve transduction efficiency and the safety of viral vectors, including retrovirus (4, 7), lentivirus (25), vesicular stomatitis virus (VSV) (29), and baculovirus (17, 42). Pseudotype retroviruses and lentiviruses bearing the baculovirus envelope protein GP64 of Autographa californica nucleopolyhedrosis virus (AcNPV) have been shown to exhibit efficient gene transduction into a wide variety of cells with a lower cytotoxicity compared to those bearing the VSV envelope protein G (VSVG), which is commonly used for pseudotyping (18, 32, 35, 36).However, a drawback of gene transduction by viral vectors is that human sera inactivate the vectors (11, 40). Complement is a major element of the innate immune response and serves to link innate and adaptive immunity (8). Complement activation can occur via classical, lectin, and alternative pathways (2, 8). All pathways invoke several responses, such as virus opsonization, virolysis, anaphylatoxin, and chemotaxin production, as well as others (2, 8). VSV and baculovirus are inactivated by human sera via the classical pathway (1, 11). Because complement activation also induces potential damage to host cells, the complement system is tightly regulated by the complement regulatory proteins (CRPs), including CD55/decay-accelerating factor (DAF), CD46/membrane cofactor protein (MCP), and CD59 (2, 8, 15). DAF and CD46 inhibit activation of C3/C5-converting enzymes, which regulate the activation of classical and alternative pathways, whereas CD59 regulates the assembly of the membrane attack complex (2, 8, 15).Viral vectors can be manipulated to confer resistance to the complement inactivation. Human immunodeficiency virus (HIV) is known to develop resistance to human complement through the incorporation of DAF, CD46, and CD59 to the viral particles (22, 30, 31, 38). Moloney murine leukemia virus vectors produced in HT1080 cells are resistant to complement inactivation (5). Baculovirus and lentivirus vectors bearing DAF or the fusion protein between the functional domains of human DAF and the GP64 were resistant to complement inactivation (9, 13). It has been shown that lentivirus vectors pseudotyped with the GP64 are more resistant to inactivation in the sera of mice and rats (14, 32) and are capable of executing longer expression of the transgenes in nasal epithelia compared to those pseudotyped with the VSVG (35, 36). However, the precise mechanisms underlying the resistance to complement inactivation by pseudotyping of the GP64 is not known.To clarify the molecular mechanisms underlying the resistance of the viral vectors pseudotyped with the GP64 to the complement inactivation, we produced pseudotype and recombinant VSVs bearing the GP64. The recombinant VSVs carrying the gp64 gene generated in human cells but not in insect cells exhibited incorporation of human DAF on the viral particles and were resistant to the complement inactivation. Furthermore, production of the gp64 pseudotype VSV in the DAF knockdown human cells impaired serum resistance, whereas production of the gp64 recombinant VSV in the CHO cell lines stably expressing human DAF and the recombinant baculoviruses in the insect cells stably expressing human DAF or encoding the DAF gene in the genome conferred resistance to the complement inactivation. These results suggest that DAF incorporation into viral particles bearing baculovirus GP64 confers resistance to serum inactivation.  相似文献   

7.
Human immunodeficiency virus (HIV-1) entry into cells is mediated by a trimeric complex consisting of noncovalently associated gp120 (exterior) and gp41 (transmembrane) envelope glycoproteins. The binding of gp120 to receptors on the target cell alters the gp120-gp41 relationship and activates the membrane-fusing capacity of gp41. Interaction of gp120 with the primary receptor, CD4, results in the exposure of the gp120 third variable (V3) loop, which contributes to binding the CCR5 or CXCR4 chemokine receptors. We show here that insertions in the V3 stem or polar substitutions in a conserved hydrophobic patch near the V3 tip result in decreased gp120-gp41 association (in the unliganded state) and decreased chemokine receptor binding (in the CD4-bound state). Subunit association and syncytium-forming ability of the envelope glycoproteins from primary HIV-1 isolates were disrupted more by V3 changes than those of laboratory-adapted HIV-1 envelope glycoproteins. Changes in the gp120 β2, β19, β20, and β21 strands, which evidence suggests are proximal to the V3 loop in unliganded gp120, also resulted in decreased gp120-gp41 association. Thus, a gp120 element composed of the V3 loop and adjacent beta strands contributes to quaternary interactions that stabilize the unliganded trimer. CD4 binding dismantles this element, altering the gp120-gp41 relationship and rendering the hydrophobic patch in the V3 tip available for chemokine receptor binding.The entry of human immunodeficiency virus type 1 (HIV-1) is mediated by the viral envelope glycoproteins (9, 79). The HIV-1 envelope glycoproteins are synthesized as an ∼850-amino acid precursor, which trimerizes and is posttranslationally modified by carbohydrates to create a 160-kDa glycoprotein (gp160). The gp160 envelope glycoprotein precursor is proteolytically processed in the Golgi apparatus, resulting in a gp120 exterior envelope glycoprotein and a gp41 transmembrane envelope glycoprotein (16, 17, 66, 76). In the mature HIV-1 envelope glycoprotein trimer, the three gp120 subunits are noncovalently bound to three membrane-anchored gp41 subunits (32).HIV-1 entry involves the binding of gp120 in a sequential fashion to CD4 and one of the chemokine receptors, CCR5 or CXCR4 (1, 8, 15, 18, 25, 36). CD4 binding triggers the formation of an activated intermediate that is competent for binding to CCR5 or CXCR4 (29, 69, 73, 78). These chemokine receptors are G protein-coupled, 7-transmembrane segment receptors with relatively short N termini. The choice of chemokine receptors is dictated primarily by the sequence of a gp120 region, the third variable (V3) loop, that exhibits variability among HIV-1 strains and becomes exposed upon CD4 binding (4, 8, 10, 33, 37, 38, 49, 59, 75). X-ray crystal structures of CD4-bound HIV-1 gp120 have revealed that the gp120 “core” consists of a gp41-interactive inner domain, a surface-exposed and heavily glycosylated outer domain, and a conformationally flexible bridging sheet (38, 43, 79). In the CD4-bound state, the V3 loop projects 30 Å from the gp120 core, toward the chemokine receptor (38). The V3 loop in these structures consists of three elements: (i) conserved antiparallel β strands that contain a disulfide bond at the base of the loop; (ii) a conformationally flexible stem; and (iii) a conserved tip (37, 38). During the virus entry process, the base of the gp120 V3 loop and elements of the bridging sheet interact with the CCR5 N terminus, which is acidic and contains sulfotyrosine residues (12-14, 23, 24). Sulfotyrosine 14 of CCR5 is thought to insert into a highly conserved pocket near the V3 base, driving further conformational rearrangements that result in the rigidification of the V3 stem (37). The conserved β-turn at the tip of the V3 loop, along with some residues in the V3 stem, is believed to bind the “body” of CCR5, i.e., the extracellular loops and membrane-spanning helices. CCR5 binding is thought to induce further conformational changes in the HIV-1 envelope glycoproteins, leading to the fusion of the viral and target cell membranes by the gp41 transmembrane envelope glycoproteins.CCR5 binding involves two points of contact with the gp120 V3 loop: (i) the CCR5 N terminus with the V3 base and (ii) the CCR5 body with the V3 tip and distal stem (12-14, 23, 24, 37, 38). The intervening V3 stem can tolerate greater conformational and sequence variation, features that might decrease HIV-1 susceptibility to host antibodies (30). Despite amino acid variation, the length of the V3 loop is well conserved among naturally occurring group M (major group) HIV-1 strains (30, 42). This conserved length may be important for aligning the two CCR5-binding elements of the V3 loop. In addition to allowing optimal CCR5 binding, the conserved V3 length and orientation may be important for CCR5 binding to exert effects on the conformation of the HIV-1 envelope glycoproteins. We examine here the consequences of introducing extra amino acid residues into the V3 stem. The residues were introduced either into both strands of the V3 loop, attempting to preserve the symmetry of the structure, or into one of the strands, thereby kinking the loop. The effects of these changes on assembly, stability, receptor binding, and the membrane-fusing capacity of the HIV-1 envelope glycoproteins were assessed. In addition to effects on chemokine receptor binding, unexpected disruption of gp120-gp41 association was observed. Further investigation revealed a conserved patch in the tip of the V3 loop that is important for the association of gp120 with the trimeric envelope glycoprotein complex, as well as for chemokine receptor binding. Apparently, the V3 loop and adjacent gp120 structures contribute to the stability of the trimer in the unliganded HIV-1 envelope glycoproteins. These structures are known to undergo rearrangement upon CD4 binding, suggesting their involvement in receptor-induced changes in the virus entry process.  相似文献   

8.
9.
10.
The early steps of the hepatitis B virus (HBV) life cycle are still poorly understood. Indeed, neither the virus receptor at the cell surface nor the mechanism by which nucleocapsids are delivered to the cytosol of infected cells has been identified. Extensive mutagenesis studies in pre-S1, pre-S2, and most of the S domain of envelope proteins revealed the presence of two regions essential for HBV infectivity: the 77 first residues of the pre-S1 domain and a conformational motif in the antigenic loop of the S domain. In addition, at the N-terminal extremity of the S domain, a putative fusion peptide, partially overlapping the first transmembrane (TM1) domain and preceded by a PEST sequence likely containing several proteolytic cleavage sites, was identified. Since no mutational analysis of these two motifs potentially implicated in the fusion process was performed, we decided to investigate the ability of viruses bearing contiguous deletions or substitutions in the putative fusion peptide and PEST sequence to infect HepaRG cells. By introducing the mutations either in the L and M proteins or in the S protein, we demonstrated the following: (i) that in the TM1 domain of the L protein, three hydrophobic clusters of four residues were necessary for infectivity; (ii) that the same clusters were critical for S protein expression; and, finally, (iii) that the PEST sequence was dispensable for both assembly and infection processes.The hepatitis B virus (HBV) is the main human pathogen responsible for severe hepatic diseases like cirrhosis and hepatocellular carcinoma. Even though infection can be prevented by immunization with an efficient vaccine, about 2 billion people have been infected worldwide, resulting in 350 million chronic carriers that are prone to develop liver diseases (56). Current treatments consist either of the use of interferon α, which modulates antiviral defenses and controls infection in 30 to 40% of cases, or of the use of viral polymerase inhibitors that allow a stronger response to treatment but require long-term utilization and frequently lead to the outcome of resistant viruses (34, 55). A better understanding of the virus life cycle, and particularly of the mechanism by which the virus enters the cell, could provide background for therapeutics that inhibit the early steps of infection, as recently illustrated with the HBV pre-S1-derived entry inhibitor (25, 45).HBV belongs to the Hepadnaviridae family whose members infect different species. All viruses of this family share common properties. The capsid containing a partially double-stranded circular DNA genome is surrounded by a lipid envelope, in which two (in avihepadnaviruses infecting birds) or three (in orthohepadnaviruses infecting mammals) envelope proteins are embedded. A single open reading frame bearing several translation initiation sites encodes these surface proteins. Thus, the HBV envelope contains three proteins: S, M, and L that share the same C-terminal extremity corresponding to the small S protein that is crucial for virus assembly (7, 8, 46) and infectivity (1, 31, 53). These proteins are synthesized in the endoplasmic reticulum (ER), assembled, and secreted as particles through the Golgi apparatus (15, 42). The current model for the transmembrane structure of the S domain implies the luminal exposition of both N- and C-terminal extremities and the presence of four transmembrane (TM) domains: the TM1 and TM2 domains, both necessary for cotranslational protein integration into the ER membrane, and the TM3 and TM4 domains, located in the C-terminal third of the S domain (for a review, see reference 6). Among the four predicted TM domains, only the TM2 domain has a defined position between amino acids 80 and 98 of the S domain. The exact localization of the TM1 domain is still unclear, probably because of the relatively low hydrophobicity of its sequence, which contains polar residues and two prolines. The M protein corresponds to the S protein extended by an N-terminal domain of 55 amino acids called pre-S2. Its presence is dispensable for both assembly and infectivity (20, 21, 37). Finally, the L protein corresponds to the M protein extended by an N-terminal domain of 108 amino acids called pre-S1 (genotype D). The pre-S1 and pre-S2 domains of the L protein can be present either at the inner face of viral particles (on the cytoplasmic side of the ER), playing a crucial role in virus assembly (5, 8, 10, 11, 46), or on the outer face (on the luminal side of the ER), available for the interaction with target cells and necessary for viral infectivity (4, 14, 36). The pre-S translocation is independent from the M and S proteins and is driven by the L protein TM2 domain (33). Finally, HBV surface proteins are not only incorporated into virion envelopes but also spontaneously bud from ER-Golgi intermediate compartment membranes (30, 43) to form empty subviral particles (SVPs) that are released from the cell by secretion (8, 40).One approach to decipher viral entry is to interfere with the function of envelope proteins. Thus, by a mutagenesis approach, two envelope protein domains crucial for HBV infectivity have already been identified: (i) the 77 first amino acids of the pre-S1 domain (4, 36) including the myristic acid at the N-terminal extremity (9, 27) and (ii) possibly a cysteine motif in the luminal loop of the S domain (1, 31). In addition, a putative fusion peptide has been identified at the N-terminal extremity of the S domain due to its sequence homology with other viral fusion peptides (50). This sequence, either N-terminal in the S protein or internal in the L and M proteins, is conserved among the Hepadnaviridae family and shares common structural and functional properties with other fusion peptides (49, 50). Finally, a PEST sequence likely containing several proteolytic cleavage sites has been identified in the L and M proteins upstream of the TM1 domain (39). A cleavage within this sequence could activate the fusion peptide.In this study, we investigated whether the putative fusion peptide and the PEST sequence were necessary for the infection process. For this purpose, we constructed a set of mutant viruses bearing contiguous deletions in these regions and determined their infectivity using an in vitro infection model based on HepaRG cells (28). The introduction of mutations either in the L and M proteins or in only the S protein allowed us to demonstrate that, in the TM1 domain of L protein, three hydrophobic clusters not essential for viral assembly were crucial for HBV infectivity while their presence in the S protein was critical for envelope protein expression. In addition, we showed that the PEST sequence was clearly dispensable for both assembly and infection processes.  相似文献   

11.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

12.
The chlorosome envelope of Chlorobaculum tepidum contains 10 proteins that belong to four structural motif families. A previous mutational study (N.-U. Frigaard, H. Li, K. J. Milks, and D. A. Bryant, J. Bacteriol. 186:646-653, 2004) suggested that some of these proteins might have redundant functions. Six multilocus mutants were constructed to test the effects of eliminating the proteins of the CsmC/CsmD and CsmB/CsmF motif families, and the resulting strains were characterized physiologically and biochemically. Mutants lacking all proteins of either motif family still assembled functional chlorosomes, and as measured by growth rates of the mutant strains, light harvesting was affected only at the lowest light intensities tested (9 and 32 μmol photons m−2 s−1). The size, composition, and biogenesis of the mutant chlorosomes differed from those of wild-type chlorosomes. Mutants lacking proteins of the CsmC/CsmD motif family produced smaller chlorosomes than did the wild type, and the Qy absorbance maximum for the bacteriochlorophyll c aggregates in these chlorosomes was strongly blueshifted. Conversely, the chlorosomes of mutants lacking proteins of the CsmB/CsmF motif family were larger than wild-type chlorosomes, and the Qy absorption for their bacteriochlorophyll c aggregates was redshifted. When CsmH was eliminated in addition to other proteins of either motif family, chlorosomes had smaller diameters. These data show that the chlorosome envelope proteins of the CsmB/CsmF and CsmC/CsmD families play important roles in determining chlorosome size as well as the assembly and supramolecular organization of the bacteriochlorophyll c aggregates within the chlorosome.Green sulfur bacteria (GSB; phylum Chlorobi) are obligate photolithoautotrophs that utilize chlorosomes for light harvesting (2, 13). Chlorosomes additionally occur in some green-nonsulfur bacteria, also known as filamentous anoxygenic phototrophs (phylum Chloroflexi), and in a recently discovered chlorophototrophic member of the phylum Acidobacteria, “Candidatus Chloracidobacterium thermophilum” (2, 3). Chlorosomes are the largest known light-harvesting organelles and can contain up to 250,000 bacteriochlorophyll (BChl) molecules (13, 29, 30, 39). They do not have a fixed stoichiometric ratio of the major pigment, which may be BChl c, d, or e, to any protein component, and as a result they are highly variable in size, shape, and composition. In spite of this structural heterogeneity (34), the detailed molecular and supramolecular structures of the BChls in chlorosomes of Chlorobaculum tepidum were recently solved by combining systems biology, solid-state nuclear magnetic resonance (NMR), cryo-electron microscopy, and molecular modeling (22). The fundamental structural units were found to be syn-anti monomer stacks that form coaxial nanotubes, which have a 2.1-nm spacing between the adjacent BChl layers. In addition to the major BChl species, chlorosomes contain carotenoids, isoprenoid quinones, wax esters, and a small quantity of BChl a. BChl a is known to be associated with CsmA, the most highly conserved protein in chlorosomes (13).Although the structural organization of the BChl molecules in all chlorosomes may be similar (4, 22, 25, 37, 38), with the exception of CsmA, the composition and sequences of the envelope proteins of chlorosomes of the phyla Chlorobi, Chloroflexi, and Acidobacteria are not well conserved. Blankenship (1) suggested that lateral gene transfer might have been responsible for the presence of the genes for chlorosome biogenesis among some of these three groups of bacteria. However, because chlorosomes are found in each of three, early-diverging bacterial lineages that contain chlorophototrophs, two of which additionally contain homodimeric type 1 reaction centers (2, 3), it is possible that chlorosomes represent one of the earliest types of photosynthetic antennae and were present in a common ancestor of these phyla.A protein-stabilized, glycolipid envelope surrounds the chlorosome BChls, and this membrane can be considered to be an asymmetric bilayer membrane in which glycolipids form the outer leaflet and the hydrophobic tails of BChls form the inner leaflet (13, 24, 50, 53). In C. tepidum, a genetically tractable model GSB, this envelope contains 10 proteins, which are designated CsmA, CsmB, CsmC, CsmD, CsmE, CsmF, CsmH, CsmI, CsmJ, and CsmX (6-8, 14, 47, 50). The structural organization of these proteins has been studied by cross-linking and immunoblotting, which led to a model for the organization of these proteins in the chlorosome envelope (28, 50, 53). CsmA, the only protein for which any detailed structural information is available, probably binds both BChl a and carotenoids (13, 23, 31, 35, 40) and forms a large, paracrystalline array known as the “baseplate” (8, 23, 28, 35, 36, 42). The structure for apo-CsmA in an organic solvent was recently determined by NMR spectroscopy, and a model for the structural organization of CsmA in the chlorosome baseplate of C. tepidum was proposed (35, 36).Sequence comparisons suggest that the chlorosome envelope proteins can be assigned to four motif families: 1, CsmA/CsmE; 2, CsmB/CsmF (CsmH); 3, CsmC/CsmD (CsmH); and 4, CsmI/CsmJ/CsmX (48, 50). CsmA and CsmE are 49% identical and are both synthesized as precursors, which are proteolytically processed by the removal of ∼20 amino acids at their carboxy termini to generate the mature polypeptides (7, 8). CsmB and CsmF are 29% identical and 63% similar in sequence (6, 50). Moreover, the amino-terminal domain of CsmH is related in sequence to these two proteins (50). The CsmC and CsmD proteins are 26% identical and 45% similar in sequence, and these two proteins additionally share sequence similarity to the carboxyl-terminal region of CsmH. The other three chlorosome proteins (CsmI, CsmJ, and CsmX) share some sequence similarities to the precursor forms of CsmA and CsmE in their carboxyl-terminal regions, while their amino-terminal domains are obviously related to adrenodoxin-type [2Fe-2S] ferredoxins (47-50). These sequence relationships strongly imply that gene duplication and divergence have occurred among a small number of ancestral gene types, and these observations additionally suggest that some of these proteins might be functionally redundant (47, 50). This view was supported by mutational studies that showed that only CsmA was essential for the viability of C. tepidum. Mutants lacking any other single chlorosome protein still assembled functional chlorosomes that were similar in pigment composition and functionality to those of the wild type (14).Because of the possible functional redundancy of chlorosome proteins of the different motif classes, double, triple, and quadruple mutants were constructed to study the roles of the CsmC/CsmD/CsmH and CsmB/CsmF/CsmH protein motif families in chlorosome biogenesis and structure. Mutants lacking CsmI, CsmJ, and CsmX, which form the Fe/S motif family of envelope proteins, were also constructed, and these mutants will be described in detail elsewhere (27; H. Li, N.-U. Frigaard, and D. A. Bryant, unpublished data). The results presented here show that functional chlorosomes assemble in the complete absence of proteins of the CsmC/CsmD or CsmB/CsmF motif families, but the size, shape, and composition of the resulting chlorosomes are altered. The results suggest that the chlorosome envelope proteins may also influence the structural organization of the BChls in chlorosomes and thus help to define chlorosome assembly and shape.  相似文献   

13.
Tetherin (CD317/BST-2), an interferon-induced membrane protein, restricts the release of nascent retroviral particles from infected cell surfaces. While human immunodeficiency virus type 1 (HIV-1) encodes the accessory gene vpu to overcome the action of tetherin, the lineage of primate lentiviruses that gave rise to HIV-2 does not. It has been previously reported that the HIV-2 envelope glycoprotein has a Vpu-like function in promoting virus release. Here we demonstrate that the HIV-2 Rod envelope glycoprotein (HIV-2 Rod Env) is a tetherin antagonist. Expression of HIV-2 Rod Env, but not that of HIV-1 or the closely related simian immunodeficiency virus (SIV) SIVmac1A11, counteracts tetherin-mediated restriction of Vpu-defective HIV-1 in a cell-type-specific manner. This correlates with the ability of the HIV-2 Rod Env to mediate cell surface downregulation of tetherin. Antagonism requires an endocytic motif conserved across HIV/SIV lineages in the gp41 cytoplasmic tail, but specificity for tetherin is governed by extracellular determinants in the mature Env protein. Coimmunoprecipitation studies suggest an interaction between HIV-2 Rod Env and tetherin, but unlike studies with Vpu, we found no evidence of tetherin degradation. In the presence of HIV-2 Rod Env, tetherin localization is restricted to the trans-Golgi network, suggesting Env-mediated effects on tetherin trafficking sequester it from virus assembly sites on the plasma membrane. Finally, we recapitulated these observations in HIV-2-infected CD4+ T-cell lines, demonstrating that tetherin antagonism and sequestration occur at physiological levels of Env expression during virus replication.Various stages of the replication cycle of primate lentiviruses can be targeted by host antiviral restriction factors (reviewed in reference 49). In addition to the well-characterized antiviral effects of members of the APOBEC3 family of cytidine deaminases, particularly APOBEC3G and -3F, and species-specific variants of tripartite motif family 5α, the release of nascent retroviral particles has recently been shown to be a target for a novel restriction factor, tetherin (CD317/bone marrow stromal cell antigen 2 [BST-2]) (31, 46). Tetherin is an interferon-inducible gene that was originally shown to impart a restriction on the release of mutants of human immunodeficiency virus type 1 (HIV-1) that lack a vpu gene (31, 46). In tetherin-positive cells, mature Vpu-defective HIV-1 particles are retained on the cell surface, linked to the plasma membrane (PM) and each other via protease-sensitive tethers, and can be subsequently endocytosed and accumulate in late endosomes (30, 31). Tetherin is not HIV specific and restricts the release of virus-like particles derived from all retroviruses tested (18), as well as those of filoviruses and arenaviruses (18, 19, 39).Tetherin is a small (181-amino-acid) type II membrane protein with an unusual topology that exists mainly as a disulfide-linked dimer (34). It consists of an N-terminal cytoplasmic tail, a transmembrane anchor, an extracellular domain that includes three cysteine residues important for dimerization, a putative coiled-coil, and finally a glycophosphatidyinosityl-linked lipid anchor (22) that is essential for restriction (31). Tetherin localizes to retroviral assembly sites on the PM (18, 31), and this unusual structure is highly suggestive that tetherin restricts virion release by incorporation into the viral membrane and cross-linking virions to cells. Such a mechanism would make tetherin a powerful antiviral effector that can target an obligate part of most, if not all, enveloped virus assembly strategies. Moreover, since tetherin restriction has no specific requirement for virus protein sequences, to avoid its action, mammalian viruses have evolved to encode several distinct countermeasures that specifically inhibit tetherin''s antiviral function.The Vpu accessory protein antagonizes tetherin-mediated restriction of HIV-1 (31, 46). In the presence of Vpu, tetherin is downregulated from the cell surface (2, 46) and is targeted for degradation (10, 13, 14), although whether these processes are required for antagonism of tetherin function is unclear (27). HIV-1 Vpu displays a distinct species specificity in that it is unable to target tetherin orthologues from rhesus macaques or African green monkeys (14, 25). This differential sensitivity maps to the tetherin transmembrane domain, particularly residues that are predicted to have been under high positive selection pressure during primate evolution (14, 16, 25). This suggests that tetherin evolution may have been driven in part by viral countermeasures like Vpu. Vpu, however, is only encoded by HIV-1 and its direct simian immunodeficiency virus (SIV) lineage precursors. The majority of SIVs, including the SIVsm, the progenitor of both HIV-2 and SIVmac, do not encode a Vpu protein (21). In some of these SIVs, tetherin antagonism was recently shown to map to the nef gene (16, 51). SIV Nef proteins, however, are generally ineffective against human tetherin because they target a (G/D)DIWK motif that was deleted from the human tetherin cytoplasmic tail sometime after the divergence of humans and chimpanzees (51). This raises the question of how HIV-2 is able to overcome human tetherin, as recent data show chronically HIV-2-infected CEM T cells have reduced tetherin levels on their surface (10).Interestingly, it has long been known that the envelope glycoprotein of certain HIV-2 isolates can stimulate the release of Vpu-defective HIV-1 virions from cells we now know to be tetherin positive (5, 6, 43). HIV and SIV Envs form trimeric spikes of dimers of the surface subunit (SU-gp105 in HIV-2/SIVmac and gp120 in HIV-1) that bind CD4 and the chemokine coreceptor and gp41 (the transmembrane [TM] subunit that facilitates fusion with and entry into the target cell). Envelope precursors (gp140 or gp160) are synthesized in the endoplasmic reticulum, where they become glycosylated and are exported to the surface via the secretory pathway (8). During transit through the Golgi apparatus and possibly in endosomal compartments, the immature precursors are cleaved by furin-like proteases to form mature spikes (15, 29). Multiple endocytosis motifs in the gp41 cytoplasmic tail lead to only minor quantities of Env being exposed at the cell surface at any given time (7, 40). Recent data demonstrated that the conserved GYxxθ motif, a binding site for the clathrin adaptor protein AP-2 (3), in the membrane-proximal region of HIV-2 gp41 is required to promote Vpu-defective HIV-1 release from HeLa cells (1, 32). Based on experiments with HIV-1/HIV-2 chimeric envelopes, an additional requirement in the extracellular component was suggested (1). In this study we set out to examine the Vpu-like activity of HIV-2 envelope in light of the discovery of tetherin. We demonstrate that the HIV-2 Env is a tetherin antagonist, and we provide mechanistic insight into the basis of this antagonism.  相似文献   

14.
Lyssavirus assembly depends on the matrix protein (M). We compared lyssavirus M proteins from different genotypes for their ability to support assembly and egress of genotype 1 rabies virus (RABV). Transcomplementation of M-deficient RABV with M from European bat lyssavirus (EBLV) types 1 and 2 reduced the release of infectious virus. Stable introduction of the heterogenotypic M proteins into RABV led to chimeric viruses with reduced virus release and intracellular accumulation of virus genomes. Although the chimeras indicated genotype-specific evolution of M, rapid selection of a compensatory mutant suggested conserved mechanisms of lyssavirus assembly and the requirement for only few adaptive mutations to fit the heterogenotypic M to a RABV backbone. Whereas the compensatory mutant replicated to similar infectious titers as RABV M-expressing virus, ultrastructural analysis revealed that both nonadapted EBLV M chimeras and the compensatory mutant differed from RABV M expressing viruses in the lack of intracellular viruslike structures that are enveloped and accumulate in cisterna of the degranulated and dilated rough endoplasmic reticulum compartment. Moreover, all viruses were able to bud at the plasma membrane. Since the lack of the intracellular viruslike structures correlated with the type of M protein but not with the efficiency of virus release, we hypothesize that the M proteins of EBLV-1 and RABV differ in their target membranes for virus assembly. Although the biological function of intracellular assembly and accumulation of viruslike structures in the endoplasmic reticulum remain unclear, the observed differences could contribute to diverse host tropism or pathogenicity.Rabies virus (RABV) and rabies-related rhabdoviruses are classified to the lyssavirus genus (13). Among the lyssaviruses various genotypes have evolved with differences in host species tropism and pathogenicity. For instance, members of genotype 1 rabies viruses circulate in terrestrial mammals and in bats, whereas members of other genotypes, such as genotypes 5 and 6 (European bat lyssavirus types 1 and 2 [EBLV-1 and ELBV-2]), are mainly restricted to bats. Although EBLVs appear adapted to their bat reservoir hosts, they have retained the ability to infect terrestrial mammals including humans, as indicated by occasional infection of humans and terrestrial animals such as sheep, stone marten, or cats (9, 35, 42). These are typically dead end infections with no further spread in the new host species. Experimental infection of sheep indicated that EBLVs are less pathogenic than genotype 1 viruses in the heterologous host (1, 7, 46, 47).Although the replication potentials of various lyssaviruses appear different in a given host, the genome organization is highly conserved (11, 27), containing only five virus genes that are sequentially ordered as individual cistrons. Whereas the viral nucleoprotein N, phosphoprotein P, and large polymerase L are essential for RNA synthesis (8), the envelope components matrix protein M and glycoprotein G are required for virus release and virus infectivity, respectively (31, 32).Based on receptor binding and the apoptosis-inducing properties of the sole RABV surface antigen G, G protein has been designated as the major pathogenicity determinant of genotype 1 lyssaviruses (15, 33, 34). However, several examples for G-independent mechanisms of virus-cell interaction exist (2, 3, 6, 30, 41, 44, 45), some of which have been shown to be involved in escape from antiviral innate immune responses and to be important for virus replication in vivo.One viral protein that has been noticed as an important RABV pathogenicity determinant is the matrix protein M (12, 39). M is essential for virus assembly and release (32) and is able to support virus budding even in the absence of G (31). In addition, RABV M regulates viral RNA synthesis (14, 17) and has been described to contribute to TRAIL-dependent induction of apoptosis (25) and to mitochondrial dysfunction (18). Recently, the X-ray crystal structure analysis of M of genotype 2 Lagos bat lyssavirus demonstrated a high conservation of M structures within the rhabdovirus family, despite the lack of sequence conservation between the lyssavirus and vesiculovirus genera (19).With 92.3% identity, the amino acid sequences of lyssavirus M proteins are highly conserved within the lyssavirus genus (11) but also contain differences that either may reflect intragenotypic coevolution with other virus proteins or divergence caused by different host adaptation and/or pathogenicity. To analyze the compatibility of M proteins from bat lyssaviruses in a rabies virus genetic background, we tested the ability of EBLV M proteins to support RABV replication after transient complementation and in chimeric viruses. Replacement of RABV M with EBLV-1 M resulted in a strong decrease in infectious virus production, suggesting the presence of a rather high degree of genotype-specific constraints. Surprisingly, recovery of efficiently replicating virus was achieved after only three passages, indicating that minor changes in M or M-interacting proteins may restore full budding activity in the genotype 1 context. Moreover, none of the chimeric viruses, neither the inefficiently released chimeric viruses nor the efficiently released passaged mutant, was able to support intracytoplasmic budding of viruslike structures at membranes of the rough endoplasmic reticulum (rER). Intracellular accumulation of enveloped viruslike structures in the rER is a common phenomenon of several rabies and rabies-related viruses (36, 37) that was obviously lost in the chimeric viruses. The lack of intracellular virus assembly strongly suggests that EBLV-1 and RABV M differ in their cellular target membranes for virus assembly and budding.  相似文献   

15.
Binding to the primary receptor CD4 induces conformational changes in the human immunodeficiency virus type 1 (HIV-1) gp120 envelope glycoprotein that allow binding to the coreceptor (CCR5 or CXCR4) and ultimately trigger viral membrane-cell membrane fusion mediated by the gp41 transmembrane envelope glycoprotein. Here we report the derivation of an HIV-1 gp120 variant, H66N, that confers envelope glycoprotein resistance to temperature extremes. The H66N change decreases the spontaneous sampling of the CD4-bound conformation by the HIV-1 envelope glycoproteins, thus diminishing CD4-independent infection. The H66N change also stabilizes the HIV-1 envelope glycoprotein complex once the CD4-bound state is achieved, decreasing the probability of CD4-induced inactivation and revealing the enhancing effects of soluble CD4 binding on HIV-1 infection. In the CD4-bound conformation, the highly conserved histidine 66 is located between the receptor-binding and gp41-interactive surfaces of gp120. Thus, a single amino acid change in this strategically positioned gp120 inner domain residue influences the propensity of the HIV-1 envelope glycoproteins to negotiate conformational transitions to and from the CD4-bound state.Human immunodeficiency virus type 1 (HIV-1), the cause of AIDS (6, 29, 66), infects target cells by direct fusion of the viral and target cell membranes. The viral fusion complex is composed of gp120 and gp41 envelope glycoproteins, which are organized into trimeric spikes on the surface of the virus (10, 51, 89). Membrane fusion is initiated by direct binding of gp120 to the CD4 receptor on target cells (17, 41, 53). CD4 binding creates a second binding site on gp120 for the chemokine receptors CCR5 and CXCR4, which serve as coreceptors (3, 12, 19, 23, 25). Coreceptor binding is thought to lead to further conformational changes in the HIV-1 envelope glycoproteins that facilitate the fusion of viral and cell membranes. The formation of an energetically stable six-helix bundle by the gp41 ectodomain contributes to the membrane fusion event (9, 10, 79, 89, 90).The energy required for viral membrane-cell membrane fusion derives from the sequential transitions that the HIV-1 envelope glycoproteins undergo, from the high-energy unliganded state to the low-energy six-helix bundle. The graded transitions down this energetic slope are initially triggered by CD4 binding (17). The interaction of HIV-1 gp120 with CD4 is accompanied by an unusually large change in entropy, which is thought to indicate the introduction of order into the conformationally flexible unliganded gp120 glycoprotein (61). In the CD4-bound state, gp120 is capable of binding CCR5 with high affinity; moreover, CD4 binding alters the quaternary structure of the envelope glycoprotein complex, resulting in the exposure of gp41 ectodomain segments (27, 45, 77, 92). The stability of the intermediate state induced by CD4 binding depends upon several variables, including the virus (HIV-1 versus HIV-2/simian immunodeficiency virus [SIV]), the temperature, and the nature of the CD4 ligand (CD4 on a target cell membrane versus soluble forms of CD4 [sCD4]) (30, 73). For HIV-1 exposed to sCD4, if CCR5 binding occurs within a given period of time, progression along the entry pathway continues. If CCR5 binding is impeded or delayed, the CD4-bound envelope glycoprotein complex decays into inactive states (30). In extreme cases, the binding of sCD4 to the HIV-1 envelope glycoproteins induces the shedding of gp120 from the envelope glycoprotein trimer (31, 56, 58). Thus, sCD4 generally inhibits HIV-1 infection by triggering inactivation events, in addition to competing with CD4 anchored in the target cell membrane (63).HIV-1 isolates vary in sensitivity to sCD4, due in some cases to a low affinity of the envelope glycoprotein trimer for CD4 and in other cases to differences in propensity to undergo inactivating conformational transitions following CD4 binding (30). HIV-1 isolates that have been passaged extensively in T-cell lines (the tissue culture laboratory-adapted [TCLA] isolates) exhibit lower requirements for CD4 than primary HIV-1 isolates (16, 63, 82). TCLA viruses bind sCD4 efficiently and are generally sensitive to neutralization compared with primary HIV-1 isolates. Differences in sCD4 sensitivity between primary and TCLA HIV-1 strains have been mapped to the major variable loops (V1/V2 and V3) of the gp120 glycoprotein (34, 42, 62, 81). Sensitivity to sCD4 has been shown to be independent of envelope glycoprotein spike density or the intrinsic stability of the envelope glycoprotein complex (30, 35).In general, HIV-1 isolates are more sensitive to sCD4 neutralization than HIV-2 or SIV isolates (4, 14, 73). The relative resistance of SIV to sCD4 neutralization can in some cases be explained by a reduced affinity of the envelope glycoprotein trimer for sCD4 (57); however, at least some SIV isolates exhibit sCD4-induced activation of entry into CD4-negative, CCR5-expressing target cells that lasts for several hours after exposure to sCD4 (73). Thus, for some primate immunodeficiency virus envelope glycoproteins, activated intermediates in the CD4-bound conformation can be quite stable.The HIV-1 envelope glycoprotein elements important for receptor binding, subunit interaction, and membrane fusion are well conserved among different viral strains (71, 91). Thus, these elements represent potential targets for inhibitors of HIV-1 entry. Understanding the structure and longevity of the envelope glycoprotein intermediates along the virus entry pathway is relevant to attempts at inhibition. For example, peptides that target the heptad repeat 1 region of gp41 exhibit major differences in potency against HIV-1 strains related to efficiency of chemokine receptor binding (20, 21), which is thought to promote the conformational transition to the next step in the virus entry cascade. The determinants of the duration of exposure of targetable HIV-1 envelope glycoprotein elements during the entry process are undefined.To study envelope glycoprotein determinants of the movement among the distinct conformational states along the HIV-1 entry pathway, we attempted to generate HIV-1 variants that exhibit improved stability. Historically, labile viral elements have been stabilized by selecting virus to replicate under conditions, such as high temperature, that typically weaken protein-protein interactions (38, 39, 76, 102). Thus, we subjected HIV-1 to repeated incubations at temperatures between 42°C and 56°C, followed by expansion and analysis of the remaining replication-competent virus fraction. In this manner, we identified an envelope glycoprotein variant, H66N, in which histidine 66 in the gp120 N-terminal segment was altered to asparagine. The resistance of HIV-1 bearing the H66N envelope glycoproteins to changes in temperature has been reported elsewhere (37). Here, we examine the effect of the H66N change on the ability of the HIV-1 envelope glycoproteins to negotiate conformational transitions, either spontaneously or in the presence of sCD4. The H66N phenotype was studied in the context of both CD4-dependent and CD4-independent HIV-1 variants.  相似文献   

16.
17.
Lentiviruses, including human immunodeficiency virus type 1 (HIV-1), typically encode envelope glycoproteins (Env) with long cytoplasmic tails (CTs). The strong conservation of CT length in primary isolates of HIV-1 suggests that this factor plays a key role in viral replication and persistence in infected patients. However, we report here the emergence and dominance of a primary HIV-1 variant carrying a natural 20-amino-acid truncation of the CT in vivo. We demonstrated that this truncation was deleterious for viral replication in cell culture. We then identified a compensatory amino acid substitution in the matrix protein that reversed the negative effects of CT truncation. The loss or rescue of infectivity depended on the level of Env incorporation into virus particles. Interestingly, we found that a virus mutant with defective Env incorporation was able to spread by cell-to-cell transfer. The effects on viral infectivity of compensation between the CT and the matrix protein have been suggested by in vitro studies based on T-cell laboratory-adapted virus mutants, but we provide here the first demonstration of the natural occurrence of similar mechanisms in an infected patient. Our findings provide insight into the potential of HIV-1 to evolve in vivo and its ability to overcome major structural alterations.The envelope glycoprotein complex of the human immunodeficiency virus type 1 (HIV-1) is involved principally in virion attachment to target cell surfaces and in the entry process (15, 18, 27, 29, 52). Envelope glycoproteins (Env) are initially translated as a gp160 precursor glycoprotein, which is then processed during its trafficking through the secretory pathway, to yield a surface subunit gp120 noncovalently attached to a transmembrane subunit gp41. During HIV-1 assembly, Env proteins are incorporated at the surface of the viral particle as a trimeric structure consisting of three gp120/gp41 dimers (59, 62).The gp41 consists of an ectodomain, a hydrophobic transmembrane anchor, and a cytoplasmic tail (CT). Lentiviruses, including HIV-1 and simian immunodeficiency virus (SIV), are unusual in having a transmembrane subunit with much longer CTs (∼150 amino acids) than most other retroviruses (20 to 50 amino acids) (27). Early studies with T-cell laboratory-adapted HIV-1 mutants showed that the gp41 CT region played an important role in regulating Env functions, the incorporation of Env into virus particles and, consequently, viral replication (16, 21, 35, 63). The integrity of the gp41 CT thus appears to be crucial for replication in primary T cells, macrophages, and in many transformed T-cell lines (1, 44). Viral variants with truncated gp41 are rarely isolated from infected patients. One study reported the isolation of a CD4-independent variant harboring a sharply truncated CT (64). However, this atypical isolate existed as a minority variant in the original quasispecies of the patient (54). SIV variants with truncated CTs obtained in cell culture in vitro have also been shown to revert rapidly (to full-length CT) when introduced into macaques (39). These observations indicate that the long CTs of lentiviruses, such as HIV-1 and SIV, have functions specific to viral replication and persistence in vivo.Two groups of conserved sequence motifs have been identified in the gp41 CT that are likely to be involved in its functions. The first group, involved in regulating the intracellular trafficking of Env, includes a membrane-proximal tyrosine-based endocytic motif, Y712SPL, (9, 47); a diaromatic motif, Y802W803, implicated in the retrograde transport of Env to the trans-Golgi network (8), and a C-terminal dileucine motif recently identified as a second endocytic motif (7, 10, 60). We have also provided evidence for the existence of additional as-yet-unidentified signals in studies of primary HIV-1 (34). The second group of motifs consists of three structurally conserved amphipathic α-helical domains: lentivirus lytic peptides 1, 2, and 3 (LLP-1, LLP-2, and LLP-3) (11, 17, 33). LLP domains have been implicated in various functions, including Env fusogenicity and the incorporation of Env into HIV-1 particles (28, 32, 43, 45, 50, 61).Several lines of evidence suggest that Env incorporation requires direct or indirect interactions between the matrix domain of the structural protein precursor Pr55Gag (matrix) and the gp41 CT during HIV-1 assembly. This possibility was first suggested by the observation that HIV-1 Env drives the basolateral budding of Gag in polarized cells (37, 48). A direct interaction between the matrix and a glutathione S-transferase fusion protein containing Env CT was subsequently observed in vitro (13). Synthetic peptides corresponding to various domains of the gp41 CT have also been shown to interact directly with Pr55Gag molecules (26). Furthermore, effects on viral infectivity of compensation between the CT and the matrix protein have been suggested by studies based on T-cell laboratory-adapted virus mutants (19, 40, 43). Finally, the cellular protein TIP47 was recently implicated in Env incorporation, based on its ability to bind both the matrix protein and the gp41 CT (38).In a previous study describing the evolutionary dynamics of the glycan shield of HIV-1 Env, we identified a patient (patient 153) for whom the 15 env clones obtained during primary infection (early stage) encoded full-length Env, whereas the 15 env sequences from the HIV-1 present 6 years later (late stage) encoded truncated gp41 CTs (14). These late-stage sequences contained a deletion introducing an in-frame stop codon, resulting in a 20-amino-acid truncation of the Env. Note that, unlike a point mutation, this deletion cannot easily revert to the full-length form. Such a deletion affecting various known motifs of the gp41 CT would be expected to impair viral replication. However, the plasma viral load measured in patient 153 demonstrated that the virus had retained its ability to replicate.In the present study, we explored the molecular mechanisms by which a primary HIV-1 maintained its capacity to replicate efficiently in this patient and demonstrated for the first time the occurrence of matrix and Env coevolution in vivo, providing insight into the ability of HIV-1 to overcome major structural alterations.  相似文献   

18.
The native envelope (Env) spike on the surface of human immunodeficiency virus type 1 (HIV-1) is trimeric, and thus trimeric Env vaccine immunogens are currently being explored in preclinical immunogenicity studies. Key challenges have included the production and purification of biochemically homogeneous and stable trimers and the evaluation of these immunogens utilizing standardized virus panels for neutralization assays. Here we report the binding and neutralizing antibody (NAb) responses elicited by clade A (92UG037.8) and clade C (CZA97.012) Env gp140 trimer immunogens in guinea pigs. These trimers have been selected and engineered for optimal biochemical stability and have defined antigenic properties. Purified gp140 trimers with Ribi adjuvant elicited potent, cross-clade NAb responses against tier 1 viruses as well as detectable but low-titer NAb responses against select tier 2 viruses from clades A, B, and C. In particular, the clade C trimer elicited NAbs that neutralized 27%, 20%, and 47% of tier 2 viruses from clades A, B, and C, respectively. Heterologous DNA prime, protein boost as well as DNA prime, recombinant adenovirus boost regimens expressing these antigens, however, did not result in an increased magnitude or breadth of NAb responses in this system. These data demonstrate the immunogenicity of stable, homogeneous clade A and clade C gp140 trimers and exemplify the utility of standardized tier 1 and tier 2 virus panels for assessing the NAb responses of candidate HIV-1 Env immunogens.The development and evaluation of novel HIV-1 Env immunogens are critical priorities of the HIV-1 vaccine field (2, 10, 25). The major antigenic target for neutralizing antibodies (NAbs) is the trimeric Env glycoprotein on the virion surface (4, 18, 30). Monomeric gp120 immunogens have not elicited broadly reactive NAbs in animal models (5, 13, 28, 29) or humans (16, 31), and thus several groups have focused on generating trimer immunogens that better mimic the native Env spike found on virions (3, 7, 14, 15, 20, 22, 27). It has, however, proven difficult to produce stable and conformationally homogeneous Env trimers. Strategies to modify Env immunogens have therefore been explored, including the removal of the cleavage site between gp120 and gp41 (3, 7, 23, 39, 40), the incorporation of an intramolecular disulfide bond to stabilize cleaved gp120 and gp41 moieties (6), and the addition of trimerization motifs such as the T4 bacteriophage fibritin “fold-on” (Fd) domain (8, 17, 39).Preclinical evaluation of candidate Env immunogens is critical for concept testing and for the prioritization of vaccine candidates. Luciferase-based virus neutralization assays with TZM.bl cells (21, 24) have been developed as high-throughput assays that can be standardized (26). However, the optimal use of this assay requires the generation of standardized virus panels derived from multiple clades that reflect both easy-to-neutralize (tier 1) and primary isolate (tier 2) viruses (21, 24). A tiered approach for the evaluation of novel Env immunogens has been proposed, in which tier 1 viruses represent homologous vaccine strains and a small number of heterologous neutralization-sensitive viruses while tier 2 viruses provide a greater measure of neutralization breadth for the purpose of comparing immunogens (24).We screened a large panel of primary HIV-1 isolates for Env stability and identified two viruses, CZA97.012 (clade C) (32) and 92UG037.8 (clade A) (17), that yielded biochemically homogeneous and stable Env trimers with well defined and uniform antigenic properties (17). The addition of the T4 bacteriophage fibritin “fold-on” (Fd) trimerization domain further increased their yield and purity (17). In the present study, we assessed the immunogenicity of these stable clade A and clade C gp140 trimers in guinea pigs. Both trimers elicited high-titer binding antibody responses and cross-clade neutralization of select tier 1 viruses as well as low-titer but detectable NAb responses against select tier 2 viruses from clades A, B, and C. These data demonstrate the immunogenicity of these stable gp140 trimers and highlight the utility of standardized virus panels in the evaluation of novel HIV-1 Env immunogens.  相似文献   

19.
An effective AIDS vaccine must control highly diverse circulating strains of human immunodeficiency virus type 1 (HIV-1). Among HIV-1 gene products, the envelope (Env) protein contains variable as well as conserved regions. In this report, an informatic approach to the design of T-cell vaccines directed to HIV-1 Env M group global sequences was tested. Synthetic Env antigens were designed to express mosaics that maximize the inclusion of common potential T-cell epitope (PTE) 9-mers and minimize the inclusion of rare epitopes likely to elicit strain-specific responses. DNA vaccines were evaluated using intracellular cytokine staining in inbred mice with a standardized panel of highly conserved 15-mer PTE peptides. One-, two-, and three-mosaic sets that increased theoretical epitope coverage were developed. The breadth and magnitude of T-cell immunity stimulated by these vaccines were compared to those for natural strain Envs; additional comparisons were performed on mutant Envs, including gp160 or gp145 with or without V regions and gp41 deletions. Among them, the two- or three-mosaic Env sets elicited the optimal CD4 and CD8 responses. These responses were most evident in CD8 T cells; the three-mosaic set elicited responses to an average of eight peptide pools, compared to two pools for a set of three natural Envs. Synthetic mosaic HIV-1 antigens can therefore induce T-cell responses with expanded breadth and may facilitate the development of effective T-cell-based HIV-1 vaccines.The development of AIDS vaccines has been advanced recently by demonstrations of increased survival and decreased viral load following vaccination with T-cell vaccines in nonhuman primate models (12, 19, 23, 26, 31, 37). Although such vaccine studies have implied that T cells may contribute to the control of viremia in the highly lethal simian immunodeficiency virus SIVmac251 challenge model, the applicability of these results in human studies remains uncertain. The major concern regarding the efficacy of human immunodeficiency virus (HIV) vaccines in humans is the extraordinary genetic diversity of the virus. The sequence similarity of HIV type 1 (HIV-1) envelope from diverse isolates within a clade can diverge by as much as 15%, and divergence between alternative clades may approach 30% (10). In addition, the diversity of the viral Gag gene product can approach similar levels, particularly in p17 and p15, which are much more diverse than p24 (6), although Gag does not have the extreme localized diversity seen in the highly variable regions of Env (6, 10). While the approach to viral diversity has been addressed in existing vaccines through the use of envelopes derived from representative viruses in the major clades, increasing knowledge about the genetic diversity of naturally occurring isolates has enabled alternative approaches that enhance population coverage of vaccine-elicited T-cell responses.Approaches under consideration include the use of central gene sequences based on ancestral, consensus, or center-of-the-tree genetic analyses (5, 10, 18, 31, 36). Such prototypes are derived by selection of the most common amino acids at each residue (10, 16, 17, 21, 25, 36), identifying the most recent common ancestor of diverging viruses in a vaccine target population (5, 10, 18, 36), or modeling the sequence at the center of the phylogenetic tree (29), respectively. Peptides based on any of these three centralized protein strategies enhanced the detection of T-cell responses in natural infection relative to the use of peptides based on natural strains; however, all three strategies behaved equivalently (7).The use of a single M group consensus/ancestral Env sequence has been shown to elicit T-cell responses with greater breadth of cross-reactivity than single natural strains in animal models (31, 36). Such central sequences do not exist in nature, and even phylogenetic ancestral reconstructions are just an approximate model of an ancestral state of the virus (8). Thus, central sequence strategies have provided evidence that various informatically derived gene products can elicit immune responses to T-cell epitopes found in diverse circulating strains, leading to the possibility of using computational strategies to design polyvalent vaccines which optimize T-cell coverage (6, 24). In this study, we have evaluated for the first time the ability of nonnatural mosaic Env immunogens (6) to elicit T-cell responses of increased cross-reactivity against epitopes represented in naturally circulating viruses in animals.Mosaic HIV-1 envelope genes were derived using an informatic approach, whereby in silico-generated recombinants of natural variants from the Los Alamos database M group Env alignment were created, scored, and selected in combination to optimize the coverage of 9-mers in the global database for a given vaccine cocktail size. While mosaic proteins are artificial constructs that do not occur in nature, they align well to natural proteins, and any short span found in mosaics will tend to be found repeatedly among natural strains (although some of the hypervariable loop regions of Env are so extremely variable that they are not repeated among circulating strains, and this necessitates bridging these regions with segments found in a single strain). In silico recombination breakpoints are constrained to create fusion points found in natural sequences. It is possible to provide increased breadth of coverage with a single mosaic, providing the maximum possible single-antigen diversity coverage for stretches of nine amino acids. Alternatively, multiple mosaics can increase the breadth of representation but have the drawback of requiring the synthesis of additional vectors for clinical use. Mosaics also preserve a natural Env-like sequence to retain normal antigen processing. Here, we have compared single-, double-, or triple-mosaic envelope antigen sets to naturally circulating strains or other derivatives for their ability to elicit immune responses of increased breadth. The data suggest that mosaic HIV-1 envelope sequences provide an approach that may be useful in the development of HIV vaccines that respond to T-cell epitopes represented in naturally circulating strains.  相似文献   

20.
Human cytomegalovirus (HCMV) depends upon a five-protein complex, gH/gL/UL128-131, to enter epithelial and endothelial cells. A separate HCMV gH/gL-containing complex, gH/gL/gO, has been described. Our prevailing model is that gH/gL/UL128-131 is required for entry into biologically important epithelial and endothelial cells and that gH/gL/gO is required for infection of fibroblasts. Genes encoding UL128-131 are rapidly mutated during laboratory propagation of HCMV on fibroblasts, apparently related to selective pressure for the fibroblast entry pathway. Arguing against this model in the accompanying paper by B. J. Ryckman et al. (J. Virol., 84:2597-2609, 2010), we describe evidence that clinical HCMV strain TR expresses a gO molecule that acts to promote endoplasmic reticulum (ER) export of gH/gL and that gO is not stably incorporated into the virus envelope. This was different from results involving fibroblast-adapted HCMV strain AD169, which incorporates gO into the virion envelope. Here, we constructed a TR gO-null mutant, TRΔgO, that replicated to low titers, spread poorly among fibroblasts, but produced normal quantities of extracellular virus particles. TRΔgO particles released from fibroblasts failed to infect fibroblasts and epithelial and endothelial cells, but the chemical fusogen polyethylene glycol (PEG) could partially overcome defects in infection. Therefore, TRΔgO is defective for entry into all three cell types. Defects in entry were explained by observations showing that TRΔgO incorporated about 5% of the quantities of gH/gL in extracellular virus particles compared with that in wild-type virions. Although TRΔgO particles could not enter cells, cell-to-cell spread involving epithelial and endothelial cells was increased relative to TR, apparently resulting from increased quantities of gH/gL/UL128-131 in virions. Together, our data suggest that TR gO acts as a chaperone to promote ER export and the incorporation of gH/gL complexes into the HCMV envelope. Moreover, these data suggest that it is gH/gL, and not gH/gL/gO, that is present in virions and is required for infection of fibroblasts and epithelial and endothelial cells. Our observations that both gH/gL and gH/gL/UL128-131 are required for entry into epithelial/endothelial cells differ from models for other beta- and gammaherpesviruses that use one of two different gH/gL complexes to enter different cells.Human cytomegalovirus (HCMV) infects a broad spectrum of cell types in vivo, including epithelial and endothelial cells, fibroblasts, monocyte-macrophages, dendritic cells, hepatocytes, neurons, glial cells, and leukocytes (6, 28, 36). Infection of this diverse spectrum of cell types contributes to the multiplicity of CMV-associated disease. HCMV infection of hepatocytes and epithelial cells in the gut and lungs following transplant immunosuppression is directly associated with CMV disease (3, 44). HCMV can be transported in the blood by monocyte-macrophages, and virus produced in these cells can infect endothelial cells, leading to virus spread into solid tissues such as the brain, liver, and lungs, etc. (16). Despite the broad spectrum of cells infected in vivo, propagation of HCMV in the laboratory is largely limited to normal human fibroblasts because other cells produce little virus. HCMV rapidly adapts to laboratory propagation in fibroblasts, losing the capacity to infect other cell types, i.e., epithelial and endothelial cells and monocyte-macrophages (9, 16, 18, 43). This adaptation to fibroblasts involves mutations in the unique long b′ (ULb′) region of the HCMV genome, which includes 22 genes (9). Targeted mutation of three of the ULb′ genes, UL128, UL130, and UL131, abolished HCMV infection of endothelial cells, transmission to leukocytes, and infection of dendritic cells (17, 18). Restoration of UL128-131 genes in HCMV laboratory strain AD169 (which cannot infect epithelial and endothelial cells) produced viruses capable of infecting these cells (18, 48). There is also evidence that the UL128-131 proteins are deleterious to HCMV replication in fibroblasts, resulting in rapid loss or mutation of one or more of the UL128-131 genes during passage in fibroblasts (2).A major step forward in understanding how the UL128-131 genes promote HCMV infection of epithelial and endothelial cells involved observations that the UL128-131 proteins assemble onto the extracellular domain of the membrane-anchored HCMV glycoprotein heterodimer gH/gL (1, 49). Antibodies to UL128, UL130, and UL131 each neutralized HCMV for infection of endothelial or epithelial cells (1, 49). All herpesviruses express gH/gL homologues and, where this has been tested, all depend upon gH/gL for replication and, more specifically, for entry into cells (14, 15, 31, 38). Indeed, we showed that the gH/gL/UL128-131 complex mediated entry into epithelial and endothelial cells (40). All five members of the gH/gL/UL128-131 complex were required for proper assembly and export from the endoplasmic reticulum (ER) and for function (39, 41). In addition, the expression of gH/gL/UL128-131, but not gH/gL or gB, in epithelial cells interfered with HCMV entry into these cells (39). This interference suggested that there are saturable gH/gL/UL128-131 receptors present on epithelial cells, molecules that HCMV uses for entry. There was no interference in fibroblasts expressing gH/gL/UL128-131, although some interference was observed with gH/gL (39). As noted above, gH/gL/UL128-131 plays no obvious role in entry into fibroblasts and, in fact, appears to be deleterious in this respect (2, 18, 40).HCMV also expresses a second gH/gL complex, as follows: gH/gL/gO (20, 21, 22, 30, 48). Fibroblast-adapted HCMV strain AD169 expresses a gO protein that is a 110- to 125-kDa glycoprotein (21). Pulse-chase studies suggest that gH/gL assembles first in the ER before binding and forming disulfide links with gO (21, 22). The 220-kDa immature gH/gL/gO complex is transported from the ER to the Golgi apparatus and increases in size to ∼280 to 300 kDa before incorporation into the virion envelope (21). gH/gL/gO complexes are apparently distinct from gH/gL/UL128-131 complexes because gO-specific antibodies do not detect complexes containing either UL128 or UL130 and UL128-specific antibodies do not precipitate gO (49). Towne and AD169 gO-null mutant laboratory strains can produce small plaques on fibroblasts, leading to the conclusion that gO is not essential. However, the AD169 and Towne mutants produced ∼1,000-fold less infectious virus than wild-type HCMV (14, 19), which might also be interpreted to mean that gO is very important or even essential for replication. Thus, the prevailing model has been that wild-type HCMV particles contain the following two gH/gL complexes: gH/gL/gO, which promotes infection of fibroblasts, and gH/gL/UL128-131, which promotes entry into epithelial and endothelial cells. Supporting this model, there are two different entry mechanisms, as follows: HCMV enters fibroblasts by fusion at the plasma membrane at neutral pH (12), whereas entry into epithelial and endothelial cells involves endocytosis and a low pH-dependent fusion with endosomes (40). This model of HCMV entry parallels models for Epstein-Barr virus (EBV) entry that use gH/gL to enter epithelial cells and gH/gL/gp42 to enter B cells (24). Similarly, HHV-6 uses gH/gL/gO and gH/gL/gQ, which bind to different receptors (33).Many of the studies of gH/gL/gO have involved the fibroblast-adapted HCMV strain AD169, which fails to express UL131 and assemble gH/gL/UL128-131 or AD169 recombinants in which UL131 expression was restored (20, 21, 22, 48, 49). It seemed possible that the adaptation of AD169 to long-term passage in fibroblasts might also involve alterations in gO. HCMV gO is unusually variable (15 to 25% amino acid differences) among different HCMV strains compared with other viral genes (13, 34, 35, 37, 46). In recent studies, Jiang et al. (26) described a gO-null mutant derived from the HCMV strain TB40/E, a strain that can infect endothelial cells following extensive passage on these cells. The TB40/E gO-null mutant spread poorly on fibroblasts compared with wild-type TB40/E, and there was little infectious virus detected in fibroblast culture supernatants. However, the few TB40/E gO-null mutant particles produced by fibroblasts that could initiate infection of endothelial cells were able to spread to form normal-sized plaques on endothelial cells. These results further supported the model for which gH/gL/gO is required for infection of fibroblasts but not for epithelial/endothelial cells. Those authors also concluded that gO is important for the assembly of enveloped particles in fibroblasts, based on observations of few infectious virus particles in supernatants and cytoplasmic accumulation of unenveloped capsids (26).Our studies of gH/gL/UL128-131 have involved the clinical HCMV strain TR (39, 40, 41, 47). HCMV TR was originally an ocular isolate from an AIDS patient (45) and was passaged only a few times on fibroblasts before being genetically frozen in the form of a bacterial artificial chromosome (BAC) (34, 40). HCMV TR infects epithelial and endothelial cells (40) and monocyte-macrophages (D. Streblow and J. Nelson, unpublished results) well. In the accompanying paper (42), we characterized the biochemistry and intracellular trafficking of TR gO. TR gO expressed either in TR-infected cells or by using adenovirus vectors (expressed without other HCMV proteins) was largely retained in the ER. Coexpression of gO with gH/gL promoted transport of gH/gL beyond the ER. Importantly, TR gO was not found in extracellular virions. In contrast, AD169 gO was present in extracellular virus particles, as described previously (20, 21). We concluded that TR gO is a chaperone that promotes ER export of the gH/gL complex, but gO dissociates prior to incorporation into the virus envelope. Moreover, these differences highlight major differences between gO molecules expressed by fibroblast-adapted strain AD169 and low-passage TR.To extend these results and characterize how TR gO functions, whether in virus entry or virus assembly/egress, we constructed a TR gO-null mutant. TRΔgO exhibited major defects in entering fibroblasts, as evidenced by increased virus infection following treatment with the chemical fusogen polyethylene glycol (PEG). Unexpectedly, the mutant also failed to enter epithelial and endothelial cells, and again, PEG partially restored entry. Relatively normal numbers of TRΔgO particles were produced and released into cell culture supernatants, although even with PEG treatment, most of these virus particles remained defective in initiating immediate-early HCMV protein synthesis. Western blot analyses of TRΔgO extracellular particles demonstrated very low levels of gH/gL incorporated into virions, which likely explains the reduced entry of TRΔgO. However, the small amounts of gH/gL complexes that were present in TRΔgO virions were associated with increased quantities of UL130, and these TRΔgO particles spread better than wild-type HCMV on epithelial cell monolayers. Together with the results shown in the accompanying paper (42), we concluded that HCMV TR gO functions as a chaperone to promote ER export of gH/gL to HCMV assembly compartments and the incorporation of gH/gL into the virion envelope. The highly reduced quantities of gH/gL in virions are apparently responsible for the inability of HCMV to enter fibroblasts and epithelial and endothelial cells. These results suggest a modified version of our model, in which gH/gL, not gH/gL/gO, mediates entry into fibroblasts and both gH/gL and gH/gL/UL128-131 are required for entry into epithelial and endothelial cells.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号